8
Vol. 36, no. 1 Journal of Vector Ecology 213 Implications of Plasmodium parasite infected mosquitoes on an insular avifauna: the case of Socorro Island, México Jenny S. Carlson 1 , Juan E. Martínez-Gómez 2 , Anthony Cornel 3 , Claire Loiseau 1 , and Ravinder N. M. Sehgal 1 1 Department of Biology, San Francisco State University, San Francisco, CA 94132, U.S.A. 2 Instituto de Ecología, A. C., Red de Interacciones Multitróficas, Apartado Postal 63, Xalapa, Veracruz 91000, México 3 Mosquito Control and Research Laboratory, Department of Entomology, UC Davis, 9240 S. Riverbend Avenue, Parlier, CA 93648, U.S.A. Received 1 November 2010; Accepted 2 February 2011 ABSTRACT: Avian malaria (Plasmodium spp.) has been implicated in the decline of avian populations in the Hawaiian Islands and it is generally agreed that geographically isolated and immunologically naïve bird populations are particularly vulnerable to the pathogenic effects of invasive malaria parasites. In order to assess the potential disease risk of malaria to the avifauna of Socorro Island, México, we surveyed for Plasmodium isolates from 1,300 resident field-caught mosquitoes. Most of them were identified as Aedes (Ochlerotatus) taeniorhynchus  (Wiedemann, 1821), which were abundant in the salt marshes. We also collected Culex quinquefasciatus Say, 1823 close to human dwellings. Mitochondrial ND5 and COII gene sequences of Ae. taeniorhynchus were analyzed and compared to corresponding sequences of mosquitoes of the Galápagos Islands, Latin America, and the North American mainland. Aedes lineages from Socorro Island clustered most closely with a lineage from the continental U.S. Plasmodium spp. DNA was isolated from both species of mosquitoes. From 38 positive pools, we isolated 11 distinct mitochondrial Cytb lineages of Plasmodium spp. Seven of the Plasmodium lineages represent previously documented avian infective strains while four were new lineages. Our results confirm a potential risk for the spread of avian malaria and underscore the need to monitor both the mosquito and avian populations as a necessary conservation measure to protect endangered bird species on Socorro Island. Journal of Vector Ecology 36 (1): 213-220. 2011. Keyword Index: Aedes taeniorhynchus, Culex quinquefasciatus, disease risk, avian malaria, Socorro Island. INTRODUCTION Avian malaria is a vector-borne parasitic disease caused by various species of Plasmodium and is known to have adversely affected bird populations (Valkiūnas 2005). All currently known avian Plasmodium spp. rely on mosquitoes for the completion of their natural life cycles and transmission. It has been suggested that mosquitoes play a major role in limiting geographic ranges of bird populations and in the spread of avian malaria worldwide (Valkiūnas 2005). While most studies on the prevalence and diversity of avian malaria have focused on isolates from vertebrates, some more recent studies have examined strains and lineages in field-caught mosquitoes (Ejiri et al. 2008, Gager et al. 2008, Ishtiaq et al. 2008, Kim et al. 2009, Kimura et al. 2006, Njabo et al. 2009, Tompkins and Gleeson 2006). Emerging infectious agents that are novel to any geographical region continue to pose threats to naïve endemic fauna (Cunningham et al. 2003, Daszak et al. 2000, Tompkins and Poulin 2006) and thus are recognized as a cause of wildlife extinctions worldwide (Smith et al. 2006). Due to geographical isolation, island populations are oſten particularly vulnerable to introduced pathogens (e.g., Fromont et al. 2001). For example, the introduction of Culex quinquefasciatus to the Hawaiian Islands was followed by the introduction of one or more birds infected with Plasmodium relictum (Beadell et al. 2009). Because Cx. quinquefasciatus, a competent vector for Plasmodium spp., was able to proliferate through the Hawaiian archipelago, populations of naïve endemic avifauna are reported to have declined aſter exposure to avian malaria (Atkinson and Van Riper III 1991, Van Riper et al. 1986, Warner 1968). Since these devastating consequences in Hawaii, other island avian populations have been closely monitored, most notably on the Galápagos Islands (Levin et al. 2009, Padilla et al. 2004, Parker et al. 2006, Santiago-Alarcón et al. 2010). To date, three mosquito species have been reported from the Galápagos Islands range: the southern house mosquito Culex quinquefasciatus Say, 1823, the black salt marsh mosquito Aedes (Ochlerotatus) taeniorhynchus (Wiedemann, 1821), and Aedes (Stegomyia) aegypti (Linnaeus,1762) (Peck et al. 1998). All three species of mosquitoes are known vectors of disease agents affecting wildlife and could potentially contribute to endemic wildlife declines on these islands (Bataille et al. 2009a, Bataille et al. 2009b, Levin et al. 2009, Parker et al. 2006, Peck et al. 1998, Smith et al. 2006, Snell et al. 2002, Whiteman et al. 2005, Wikelski et al. 2004). We collected mosquitoes on Socorro Island to assess the disease risk to its endemic wildlife. e island is located approximately 700 km west of the port city of Manzanillo, Colima, México. With three other volcanic islands, it is part of the Revillagigedo Archipelago Biosphere Reserve

Implications of Plasmodium parasite infected mosquitoes on ...user et al. 2011 Jour… · Mosquito identification Voucher female and all male mosquitoes were pinned and identified

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Implications of Plasmodium parasite infected mosquitoes on ...user et al. 2011 Jour… · Mosquito identification Voucher female and all male mosquitoes were pinned and identified

Vol. 36, no. 1 Journal of Vector Ecology 213

Implications of Plasmodium parasite infected mosquitoes on an insular avifauna: the case of Socorro Island, México

Jenny S. Carlson1, Juan E. Martínez-Gómez2, Anthony Cornel3, Claire Loiseau1, and Ravinder N. M. Sehgal1

1Department of Biology, San Francisco State University, San Francisco, CA 94132, U.S.A.2Instituto de Ecología, A. C., Red de Interacciones Multitróficas, Apartado Postal 63, Xalapa, Veracruz 91000, México

3Mosquito Control and Research Laboratory, Department of Entomology, UC Davis, 9240 S. Riverbend Avenue, Parlier, CA 93648, U.S.A.

Received 1 November 2010; Accepted 2 February 2011

ABSTRACT: Avian malaria (Plasmodium spp.) has been implicated in the decline of avian populations in the Hawaiian Islands and it is generally agreed that geographically isolated and immunologically naïve bird populations are particularly vulnerable to the pathogenic effects of invasive malaria parasites. In order to assess the potential disease risk of malaria to the avifauna of Socorro Island, México, we surveyed for Plasmodium isolates from 1,300 resident field-caught mosquitoes. Most of them were identified as Aedes (Ochlerotatus) taeniorhynchus  (Wiedemann, 1821), which were abundant in the salt marshes. We also collected Culex quinquefasciatus Say, 1823 close to human dwellings. Mitochondrial ND5 and COII gene sequences of Ae. taeniorhynchus were analyzed and compared to corresponding sequences of mosquitoes of the Galápagos Islands, Latin America, and the North American mainland. Aedes lineages from Socorro Island clustered most closely with a lineage from the continental U.S. Plasmodium spp. DNA was isolated from both species of mosquitoes. From 38 positive pools, we isolated 11 distinct mitochondrial Cytb lineages of Plasmodium spp. Seven of the Plasmodium lineages represent previously documented avian infective strains while four were new lineages. Our results confirm a potential risk for the spread of avian malaria and underscore the need to monitor both the mosquito and avian populations as a necessary conservation measure to protect endangered bird species on Socorro Island. Journal of Vector Ecology 36 (1): 213-220. 2011.

Keyword Index: Aedes taeniorhynchus, Culex quinquefasciatus, disease risk, avian malaria, Socorro Island.

INTRODUCTION

Avian malaria is a vector-borne parasitic disease caused by various species of Plasmodium and is known to have adversely affected bird populations (Valkiūnas 2005). All currently known avian Plasmodium spp. rely on mosquitoes for the completion of their natural life cycles and transmission. It has been suggested that mosquitoes play a major role in limiting geographic ranges of bird populations and in the spread of avian malaria worldwide (Valkiūnas 2005). While most studies on the prevalence and diversity of avian malaria have focused on isolates from vertebrates, some more recent studies have examined strains and lineages in field-caught mosquitoes (Ejiri et al. 2008, Gager et al. 2008, Ishtiaq et al. 2008, Kim et al. 2009, Kimura et al. 2006, Njabo et al. 2009, Tompkins and Gleeson 2006).

Emerging infectious agents that are novel to any geographical region continue to pose threats to naïve endemic fauna (Cunningham et al. 2003, Daszak et al. 2000, Tompkins and Poulin 2006) and thus are recognized as a cause of wildlife extinctions worldwide (Smith et al. 2006). Due to geographical isolation, island populations are often particularly vulnerable to introduced pathogens (e.g., Fromont et al. 2001). For example, the introduction of Culex quinquefasciatus to the Hawaiian Islands was followed by the introduction of one or more birds infected

with Plasmodium relictum (Beadell et al. 2009). Because Cx. quinquefasciatus, a competent vector for Plasmodium spp., was able to proliferate through the Hawaiian archipelago, populations of naïve endemic avifauna are reported to have declined after exposure to avian malaria (Atkinson and Van Riper III 1991, Van Riper et al. 1986, Warner 1968). Since these devastating consequences in Hawaii, other island avian populations have been closely monitored, most notably on the Galápagos Islands (Levin et al. 2009, Padilla et al. 2004, Parker et al. 2006, Santiago-Alarcón et al. 2010). To date, three mosquito species have been reported from the Galápagos Islands range: the southern house mosquito Culex quinquefasciatus Say, 1823, the black salt marsh mosquito Aedes (Ochlerotatus) taeniorhynchus (Wiedemann, 1821), and Aedes (Stegomyia) aegypti (Linnaeus,1762) (Peck et al. 1998). All three species of mosquitoes are known vectors of disease agents affecting wildlife and could potentially contribute to endemic wildlife declines on these islands (Bataille et al. 2009a, Bataille et al. 2009b, Levin et al. 2009, Parker et al. 2006, Peck et al. 1998, Smith et al. 2006, Snell et al. 2002, Whiteman et al. 2005, Wikelski et al. 2004).

We collected mosquitoes on Socorro Island to assess the disease risk to its endemic wildlife. The island is located approximately 700 km west of the port city of Manzanillo, Colima, México. With three other volcanic islands, it is part of the Revillagigedo Archipelago Biosphere Reserve

Page 2: Implications of Plasmodium parasite infected mosquitoes on ...user et al. 2011 Jour… · Mosquito identification Voucher female and all male mosquitoes were pinned and identified

214 Journal of Vector Ecology June 2011

(SEDESOL 1994). Wildlife is depauperate on Socorro Island, with terrestrial birds predominating among a few reptiles and only introduced mammals (Brattstrom and Howell 1956). Although Ae. taeniorhynchus was reported earlier in coastal habitats (Vázquez 1960), there is no recent documentation on the distribution and diversity of the mosquito fauna on this island. Because of its location, Socorro Island shares several biogeographical attributes with the Galápagos Islands and has a high degree of biotic endemicity (Brattstrom 1990, Levin and Moran 1989).

Socorro Island has the largest concentration of endangered birds in México (Alliance for Zero Extinction 2010, Ricketts et al. 2005), including two critically endangered species, the Socorro mockingbird (Mimus graysoni) and Townsend´s shearwater (Puffinus auricularis), the endangered Socorro Parakeet (Aratinga brevipes), the extinct in the wild Socorro dove (Zenaida graysoni), and one definitive extinction, the Socorro Elf Owl (Micrathene whitneyi graysoni; Brattstrom 1990, IUCN 2006). Habitat degradation and cat predation have been implicated in the declines of these species (Martínez-Gómez and Curry 1996, Martínez-Gómez et al. 2001, Martínez-Gómez and Jacobsen 2004). However, the presence of mosquito species that are known competent vectors for an array of pathogens presents an additional threat for the endemic birds and must be studied prior to reintroduction efforts for the Socorro dove (Martínez-Gómez et al. 2010).

We hypothesized that due to high levels of human-aided transport through boats and airplanes and a comparable location in the Eastern Pacific Ocean, Socorro Island would similarly harbor mosquito species found in the Galápagos Islands. By employing a phylogenetic approach using mitochondrial DNA (mtDNA) sequence analyses

we explored the possible origins of the mosquitoes on the island and the diversity of their haemosporidian parasites. We examine for the first time avian parasite infectivity in mosquitoes found on this remote island and discuss its conservation implications.

MATERIALS AND METHODS

Sample collectionMosquitoes were collected in July, 2009 from five sites

on the island (Figure 1): Las Grutas (Site 1; N 18° 44’ 4. 8”, W 110° 56’ 48. 3”), Los Cedros (Site 2; N 18° 45’ 28. 9”, W 110° 56’ 44. 4”), Playa Blanca (Site 3; N 18° 48’ 51”, W 111° 02’ 422. 5”), Llano de Borregos (Site 4; N 18° 45’ 57. 5”, W 110° 56 ‘58. 5”), and the pier (Site 5; N 18° 44’ 41. 6”, W 110° 56’ 30. 7”) using eight Center for Disease Control (CDC) Miniature Light Traps baited with CO2 emitting packets (John W. Hock, Gainesville, FL). Traps were hung from trees approximately 1 m above ground and 6 m apart from one another and operated from dusk until dawn (approximately 12 h). Early in the mornings, collection bags were removed from the traps and the mosquitoes were immobilized by smoke fumigation and sorted by sex. Mosquitoes were then pooled into groups of 20 and held in vials containing silica beads as a desiccant to prevent DNA degradation until identification.

Mosquito identificationVoucher female and all male mosquitoes were pinned

and identified using a stereomicroscope and morphological keys (Belkin, 1962). Genitalia were removed from the few male mosquitoes collected and slide mounted according to Belkin (1962).

Figure 1. Map of Socorro Island showing the geographical distribution of study sites where mosquito traps were set. Las Grutas (Site 1), Los Cedros (Site 2), Playa Blanca (Site 3), Llano de Borregos (Site 4), and the pier (Site 5). Circles indicate sites where no mosquitoes were trapped and triangles where successful trapping occurred (Base map courtesy of Manuel Escamilla – INECOL according to the photomap prepared by Mexico´s National Institute of Statistics, Geography and Informatics - INEGI).

Page 3: Implications of Plasmodium parasite infected mosquitoes on ...user et al. 2011 Jour… · Mosquito identification Voucher female and all male mosquitoes were pinned and identified

Vol. 36, no. 1 Journal of Vector Ecology 215

Mosquito and parasite screeningThe head/thorax of each female mosquito from each

pool was severed from the abdomen following the methods described in Ishtiaq et al. (2008). DNA was extracted from both the abdomen and head/thorax pools using the Wizard SV Genomic DNA Purification kits (Promega Corporation, Madison, WI) with the modification of adding an additional 20 µl of Proteinase K prior to incubation and homogenization with a heat-sealed pipette tip (Njabo et al. 2009). All polymerase chain reactions (PCR) were carried out in a 25 μl reaction mixture containing 10-100 ng of genomic DNA (2 µg of template DNA), 10 mMTris-HCl (pH8.3), 50 mMKCl, 3.0 mMMgCl, 0.4 mM of each dNTP, 0.4 mM of each primer, 5 μl of CL buffer (Qiagen, Valencia, CA), and 0.5 units Taq (Qiagen, Valencia, CA).

For mosquito mtDNA sequence analysis, we amplified from four pools of abdomens ~654 bp of the mitochondrial cytochrome oxidase subunit 2 gene (COII) by using the thermocycler protocols and the forward primer SCTL2-J3037 from Bataille et al. (2009b) and Cook et al. (2009), along with a modified reverse primer (5’-GATTTAAGAGATCATTACTTGC-3’; Bataille et al. 2009b). We also amplified ~610 bp of the NADH dehydrogenase subunit 5 gene (ND5) using the ND5 and tRNAPhe primers and thermocycler protocols from Krzywinski et al. (2001).

The DNA extracted from head/thorax and abdomen mosquito pools was then screened for both Plasmodium and Haemoproteus parasite DNA using PCR. We amplified ~ 750 bp of the parasite mitochondrial cytochrome oxidase subunit-b gene (cytb) using two sets of primers. The first set of primers used were L15183 and H15730 (Chasar et al. 2009, Fallon et al. 2003). The second primer set consisted of the nested primers HaemNF, HaemNR2, HaemF and HaemR2 as described in (Waldenström et al. 2004). All PCR products were screened and viewed on 1.8% agarose gels stained with ethidium bromide for examination of amplification products.

SequencingPositive PCR products were bi-directionally sequenced

by Elim Biopharmaceuticals Inc., Hayward, CA. From four pools of Ae. taeniorhynchus mosquitoes, we sequenced 654 bp of the COII gene and 588 bp of the ND5 gene. All amplicons sequenced where aligned and edited using Sequencher 4.8 (GeneCodes, Ann Arbor, MI). Because both amplicons had low levels of variation, they were combined for a total of 1,242 bp to increase the resolution power for data analysis.

From mosquito abdomen and head/thorax pools that were positive for parasite DNA, we then sequenced 750 bp of the Plasmodium spp. cytb gene and edited using Sequencher 4.8 (GeneCodes, Ann Arbor, MI). Sequences were identified as Plasmodium spp. using the NCBI nucleotide Blast search and retrieved matches ranging from 97-99% in GenBank. Parasite sequences that differed by only 1 to 3 bp were considered distinct lineages (Ricklefs and Fallon 2002). Distinct lineages were then verified by

repeating an independent PCR and sequencing analysis. Sequence chromatograms were visually inspected for presence of double peaks to ensure that no pools contained multiple distinct lineages of Plasmodium spp.

All final sequences were deposited to GenBank with accession numbers HQ853668-HQ853680. And lastly, to estimate the proportion of infected individual mosquitoes from the pooled samples, the maximum likelihood estimation (MLE) was calculated using the equation MLE= 1 – (1 – number of positive pools/number of total pools)1/pool size (Gu et al. 2009). This number represents the expected proportion of infected individuals in the sample incorporating a correction for the pooling process.

Phylogenetic analysis Phylogenetic relationships were analyzed using

MrBayes version 3.1.2 (Huelsenbeck and Ronquist 2001) using the GTR+G model for both the mosquito and para-site phylogenetic hypotheses obtained from MrModelt-est (Nylander 2004). In addition to Ae. taeniorhynchus COII and ND5 sequences obtained from Socorro Island (HQ853679- HQ853680), sequences from 32 mosquitoes originating from the Galápagos Islands, Latin America, and North America (Bataille et al. 2009b) were obtained from Genbank (accession numbers FM992116-FM992321). The COII and ND5 genes were considered a single dataset since they were characterized by the same nucleotide frequencies and the best-fit nucleotide substitution model. Due to the lack of an appropriate outgroup, we constructed an unroot-ed tree. Such an approach was not possible for Cx. quinque-fasciatus because we did not have sufficient sequence data for this species.

Plasmodium spp. mitochondrial cytb sequences were analyzed from 11 distinct lineages from Socorro Island and 34 sequences that were downloaded from Genbank. Genbank Plasmodium spp. sequences were from American mainland (Kimura et al. 2006), South Pacific (Ishtiaq et al. 2008), Japanese (Ejiri et al. 2008), and African parasite samples (Beadell et al. 2009, Chasar et al. 2009). Their respective accession numbers are provided in Figure 3. Plasmodium gallinaceum was used as an outgroup lineage. Two Markov Chain Monte Carlo (MCMC) simulations were run simultaneously for ten million generations with sampling every 200 generations generating 100,000 trees. The first 25,000 trees were discarded from the sample as the “burn-in” period that accounted for 25% of the trees. The remaining trees were used to construct a majority rule consensus tree and to calculate posterior probabilities of the individual clades (Labarthe et al. 1998).

RESULTS

Mosquitoes were observed and collected at only two of the five sampling sites, Playa Blanca (Site 3) and the pier (Site 5; Figure 1). From these two sites a total of 1,300 mosquitoes were either pinned as voucher specimens (Bohart Museum, UC Davis) or pooled for DNA isolation. Two species were identified; Culex quinquefasciatus Say, 1823 and the black

Page 4: Implications of Plasmodium parasite infected mosquitoes on ...user et al. 2011 Jour… · Mosquito identification Voucher female and all male mosquitoes were pinned and identified

216 Journal of Vector Ecology June 2011

Figure 2. Bayesian phylogeny of the combined dataset of the ND5 and COII genes of Aedes taeniorhynchus. The Bayesian posterior probabilities are depicted at each node.

Figure 3. Bayesian phylogeny of 45 mitochondrial cytochrome b Plasmodium spp. lineages found in mosquitoes and one Plasmodium gallinaceum lineage as the outgroup. The Bayesian posterior probabilities are depicted at each node. Lineages obtained from Socorro Island are delineated by the name SocP and are in bold. All other Plasmodium spp. lineages are delineated by the host name in which the parasite was found, followed by the geographical region in which it was collected in parentheses, if available, and by the Genbank accession numbers.

Page 5: Implications of Plasmodium parasite infected mosquitoes on ...user et al. 2011 Jour… · Mosquito identification Voucher female and all male mosquitoes were pinned and identified

Vol. 36, no. 1 Journal of Vector Ecology 217

salt marsh mosquito Aedes (Ochlerotatus) taeniorhynchus (Wiedemann, 1821). Identification of Cx. quinquefasciatus as a member of the Culex pipiens complex was confirmed by morphology of male genitalia (DV/D index) as described by Cornel et al. (2003). Cx. quinquefasciatus was only collected at the pier (site 5) and consisted of 13 specimens, two of which were males. Ae. taeniorhynchus was by far the most abundant species and consisted of the remaining 1,287 specimens pooled or pinned, which in fact represented only a fraction of the total numbers collected of this species. We are unable to confirm potential breeding sites since no larvae were collected, but based on the location where the two mosquito species were collected (Figure 1), we can speculate where they might occur. Culex quinquefasciatus was only collected near the Naval base where some water cisterns are not properly covered, thus providing a potential breeding site for this mosquito species (Belkin et al. 1970). The pier also harbored Ae. taeniorhynchus, a mosquito known for breeding in transient pools of salt and brackish water in low-laying coastal areas (Belkin et al. 1970). Aedes taeniorhynchus was also collected at Playa Blanca, a cove surrounded by Conocarpus erectus mangrove and temporary pools of brackish water. The volcanic nature of Socorro Island results in uneven terrain that provides ample opportunity for pools to form during rains and after high tides, which Ae. taeniorhynchus may use as breeding sites.

Female mosquitoes were pooled to create 61 separate head/thorax and abdomen pools of Ae. taeniorhynchus (20 in each) and one pool each of 11 Cx. quinquefasciatus head/thorax and abdomens for DNA analysis. Just over half of the Aedes abdomen pools (37/61) tested positive for parasite DNA, yielding a maximum likelihood estimation (MLE) of 4.8%, while the single Culex abdomen pool tested positive for parasite DNA yielding an MLE of 100%. Only one of 61 Aedes head/thorax pools tested positive for parasite DNA yielding an MLE of 0.08%.

Phylogenetic relationships among the Ae. taeniorhynchus mosquitoes from Socorro Island, Galápagos Islands, and the American mainland (Figure 2) reveal five distinct clusters of Aedes lineages: I. the Atlantic coast and Gulf of México, II. the Galápagos Islands, III. the Pacific coast, IV. Latin America, and V. Socorro Island. The Aedes lineage found on Socorro Island interestingly clustered with a lineage from the Gulf of México (Lou6 from Louisiana, U.S.A.). We should point out that the sequences used for the Socorro Island Ae. taeniorhychus were derived from four pools of 20 mosquitoes in each. Each pool produced amplicons of identical sequence for both ND5 and CO II genes.

In total, 11 distinct Plasmodium Cytb lineages were isolated. No Haemoproteus spp. were detected in the mosquitoes. Phylogenetic relationships among Plasmodium spp. found in mosquitoes from Socorro Island and those found in the American mainland, South Pacific, Japan, and Africa are depicted in Figure 3. The tree shows that Aedes isolates SocP1 and SocP8 fell in clade IV and SocP7 fell into clade III. SocP11, which was found in the only Culex pool, fell in clade II. Lineage SocP1 was the most abundant

parasite in mosquitoes sampled; it was found in 25 out of 61 Aedes pools and it included the single positive thorax/head Aedes pool. The remaining seven lineages were all found in a monophyletic clade (II in Figure 3) with Plasmodium spp. found in birds of Africa.

DISCUSSION

Our study provides significant documentation of the mosquito fauna on Socorro Island. The species collected were Cx. quinquefasciatus and Ae. taeniorhynchus, both of which occur on the Galápagos Islands (Bataille et al. 2009b). Cx. quinquefasciatus, reported for the first time on Socorro Island, is a known vector for avian pox, West Nile virus (WNV), and avian malaria in Hawaii (Atkinson et al. 1995, Van Riper et al. 1986, Warner 1968). Ae. taeniorhynchus, on the other hand, has not been implicated as a competent vector for avian malaria, but is a known vector of dog heartworm (Dirofilaria immitis) in Latin America (Labarthe et al. 1998) and a competent vector for St. Louis encephalitis (Nayar et al. 1986) and West Nile viruses (Nayar et al. 1986, Turell et al. 2001). We cannot absolutely confirm that these are the only two species on Socorro Island, as we did not place traps in a systematic manner to cover every habitat on the island. Furthermore, most trapping nights were windy and not entirely conducive to optimal mosquito capture.

Distribution patterns of these two mosquito species on Socorro Island differed. Cx. quinquefasciatus were collected only at the human-inhabited site at the pier near the Mexican Naval Station. This is consistent with findings on the Galápagos Islands where Cx. quinquefasciatus was also found only near human settlements, and that they are recent introductions. As in the Galápagos Islands (Bataille et al. 2009a, Bataille et al. 2009b), rainfall on Socorro Island is low, resulting in a scarcity of fresh water sources especially on lower altitude vegetation types (e.g., Flores-Palacios et al. 2009). Ae. taeniorhynchus was widely distributed along the coast of Socorro Island and was not caught inland nor above 200 m. This differs from the findings on the Galápagos Islands where Ae. taeniorhynchus were found up to 20 km from the coast and up to an altitude of 700 m. This unusual mosquito distribution led Bataille et al. (2009b) to suggest that Galápagos Ae. taeniorhynchus have undergone adaptations in breeding and radiated out of solely salt marshes into other breeding sites. This occurred over a period of 200,000 years, an estimate for the introduction of this species into the Galápagos Islands (Bataille et al. 2009b).

On Socorro Island, Ae. taeniorhynchus shares a more similar biology to mainland mosquitoes where it is rare to be found >6 km from the coast line, and if they are seen inland, it is only to breed (Nielsen and Nielsen 1953, Provost 1951). Our phylogenetic analysis of the ND5/COII genes (Figure 2) suggests that Ae. taeniorhynchus colonized Socorro Island more recently than the Galápagos Islands, as their sequences aligned most closely with a lineage found in the Gulf of México. These mosquitoes were most likely introduced by human activity, as was surmised with Culex mosquitoes from the Galápagos (Bataille et al. 2009b), but

Page 6: Implications of Plasmodium parasite infected mosquitoes on ...user et al. 2011 Jour… · Mosquito identification Voucher female and all male mosquitoes were pinned and identified

218 Journal of Vector Ecology June 2011

there is no documentation of when this may have occurred. Our results provide evidence that several Plasmodium

species occur on Socorro Island, but whether Cx. quinquefasciatus and Ae. taeniorhynchus are the transmitting vectors of all the lineages is not clear. Most Plasmodium spp. isolates were obtained from pools of mosquito abdomens, which does not verify their competence as vectors. To confirm vector competence, infective sporozoites must at least be detected in mosquito salivary glands (Njabo et al. 2009). One Plasmodium isolate was obtained from a head/thorax of an Aedes pool that lacked a recent blood meal, which suggests that Ae. taeniorhynchus could transmit Plasmodium spp. on the island. This is the first study providing evidence that Ae. taeniorhynchus might be a competent vector for avian malaria.

The phylogenetic tree of the parasite lineages (Figure 3) found in the mosquitoes illustrates the diversity of the Plasmodium spp. on Socorro Island. All distinct lineages found on this island did not match any other Plasmodium spp. lineages found on GenBank, suggesting that the Socorro Plasmodium spp. may be unique to this island. Clade I was composed of seven distinct parasite lineages found on Socorro Island along with three Plasmodium spp. that have been described from avian populations of Africa. Although there is one endemic lizard present on the island (Urosaurus auriculatus) (Brattstrom 1990), none of the 11 Plasmodium distinct lineages fell into the clade with two known Plasmodium spp. to infect reptiles; Plasmodium mexicanum and Plasmodium chiricahuae. This indicates that the trapped mosquitoes relied on avian populations as their main source of blood meals. However, we were unable to analyze blood meals for host identities due to the lack of recently blood-fed trapped mosquitoes during our collection. Moreover, due to the low support for the SocP11 placement in Clade II (Figure 3), which is the parasite lineage isolated in the only Culex pool, we are unable to make any assumptions of how it is related to the other Plasmodium lineages.

In order to protect the endangered avifauna of Socorro Island, and to prevent further extinctions, mosquito populations should be monitored continuously and control measures should be incorporated in the management plans of this Biosphere Reserve. Although it is unclear whether Ae. taeniorhynchus is a competent vector for avian malaria, it should be considered a serious threat to the avifauna of Socorro Island. If Ae. taeniorhynchus were to move inland, threatened birds that currently reside away from the coastline might be exposed to Plasmodium spp. Dispersal patterns of native birds require further study to help design epidemiological models to estimate disease risk and spread across the island. At this time, with a combination of timely research and sound management practices, there exists an opportunity to avert the scenario presented to the Hawaiian Islands: the collapse of an insular endemic avifauna.

Acknowledgments

We offer our sincere gratitude and recognition to the Secretaría de Marina – Armada de México, in particular, to Admiral Ezequiel Ramiro Bordonave, Captain Pedro Villaciz León and the military personnel stationed at Socorro Island for their hospitality and logistical support, without which this study would not have been possible. We thank Dr. Chindi Peavey at the San Mateo County Mosquito and Vector Control District for offering technical advice and for allowing us to use mosquito traps. Dr. Víctor Rico-Gray and Dr. Vicente Hernández at the Instituto de Ecología in Xalapa (INECOL) provided valuable insights. Funding was provided by the College of Science and Engineering, San Francisco State University, the American Society of Parasitologists Willis A. Reid Student Research Grant, the San Francisco State University Instructionally Related Activities Grant, the Island Endemics Foundation, and INECOL. All research was carried out under permit (SGPA/DGVS/04005/08) issued by the Secretaría de Medio Ambiente y Recursos Naturales. Research protocols in this study comply with IACUC permit #A9-002 SFSU.

REFERENCES CITED

Alliance for Zero Extinction. 2010. 2010 AZE Update. www.zeroextinction.org.

Atkinson, C. T. and C. Van Riper III. 1991. Vectors, epizootiology, and pathogenicity of avian species of Haemoproteus (Haemosporina: Haemoproteidae). Bull. Soc. Vector Ecol. 16: 109-126.

Atkinson, C. T., K. L. Woods, R. J. Dusek, L. S. Sileo, and W. M. Iko. 1995. Wildlife disease and conservation in Hawaii: pathogenicity of avian malaria (Plasmodium relictum) in experimentally infected iiwi (Vestiaria coccinea). Parasitology 111: S59-S69.

Bataille, A., A. A. Cunningham, V. Cedeño, M. Cruz, G. Eastwood, D. M. Fonseca, C. E. Causton, R. Azuero, J. Loayza, J. D. C. Martínez, et al. 2009a. Evidence for regular ongoing introductions of mosquito disease vectors into the Galápagos Islands. Proc. R. Soc. B 276: 3769-3775.

Bataille, A., A. A. Cunningham, V. Cedeño, L. Patiño, A. Constantinou, L. D. Kramer, and S. J. Goodman. 2009b. Natural colonization and adaptation of a mosquito species in Galápagos and its implications for disease threats to endemic wildlife. Proc. Natl. Acad. Sci. USA 106: 10230-10235.

Beadell, J. S., R. Covas, C. Gebhard, F. Ishtiaq, M. Melo, B. K. Schmidt, S. L. Perkins, G. R. Graves, and R. C. Fleischer. 2009. Host associations and evolutionary relationships of avian blood parasites from West Africa. Int. J. Parasitol. 39: 257-266.

Belkin, J. N., S. J. Heinemann, and W. A. Page. 1970. Mosquito studies (Diptera, Culicidae). XXI. The Culicidae of Jamaica. Contrib. Am. Entomol. Inst. 6: 1-458.

Belkin, J. N. 1962. The Mosquitoes of the South Pacific

Page 7: Implications of Plasmodium parasite infected mosquitoes on ...user et al. 2011 Jour… · Mosquito identification Voucher female and all male mosquitoes were pinned and identified

Vol. 36, no. 1 Journal of Vector Ecology 219

(Diptera, Culicidae). University of California Press, Berkeley.

Brattstrom, B.H. 1990. Biogeography of the Islas Revillagigedo, México. J. Biogeogr. 17:177-183.

Brattstrom, B. H. and T. R. Howell. 1956. The birds of the Revilla Gigedo Islands, Mexico. Condor 58: 107-120.

Chasar, A., C. Loiseau, G. Valkiūnas, T. Iezhova, T. B. Smith, and R. N. M. Sehgal. 2009. Prevalence and diversity patterns of avian blood parasites in degraded African rainforest habitats. Mol. Ecol. 18: 4121-4133.

Cook, S., M. Diallo, A. A. Sall, A. Cooper, and E. C. Holmes. 2009. Mitochondrial markers for molecular identification of Aedes mosquitoes (Diptera: Culicidae) involved in transmission of arboviral disease in West Africa. J. Med. Entomol. 42: 19-28.

Cornel, A. J., R. D. McAbee, J. Rasgon, M. A. Stanich, T. W. Scott, and M. Coetzee. 2003. Differences in extent of genetic introgression between sympatric Culex pipiens and Culex quinquefasciatus (Diptera: Culicidae) in California and South Africa. J. Med. Entomol. 40: 36-51.

Cunningham, A. A., P. Daszak, and J. P. Rodríguez. 2003. Pathogen pollution: defining a parasitological threat to biodiversity conservation. J. Parasitol. 89: S78-S83.

Daszak, P., A. A. Cunningham, and A. D. Hyatt. 2000. Emerging infectious diseases of wildlife threats to biodiversity and human health. Science 287: 443-449.

Ejiri, H., Y. Sato, E. Sasaki, D. Sumiyama, Y. Tsuda, K. Sawabe, S. Matsui, S. Horie, K. Akatani, and M. Takagi. 2008. Detection of avian Plasmodium spp. DNA sequences from mosquitoes captured in Minami Daito Island of Japan. J. Vet. Med. Sci. 70: 1205-1210.

Fallon, S.M., E. Bermingham, and R.E. Ricklefs. 2003. Island and taxon effects in parasitism revisited: avian malaria in the Lesser Antilles. Evolution 57: 606-615.

Flores-Palacios, A., J.E. Martínez-Gómez, and R.L. Curry. 2009. La vegetación de IslaSocorro, Archipiélago de Revillagigedo, México. B. Soc. Bot. Mex. 84: 13-23.

Fromont, E., L. Morvilliers, M. Artois, and D. Pontier. 2001. Parasite richness and abundance in insular and mainland feral cats: insularity or density? Parasitology 123: 143-151.

Gager, A. B., J. Del Rosario Loaiza, D. C. Dearborn, and E. Bermingham. 2008. Do mosquitoes filter the access of Plasmodium cytochrome b lineages to an avian host? Mol. Ecol. 17: 2552-2561.

Gu, W., R. Lampman, and R.J. Novak. 2009. Problems in estimating mosquito infection rates using minimum infection rate. J. Med. Entomol. 40: 595-596.

Huelsenbeck, J.P. and F. Ronquist. 2001. MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 754-755.

IUCN. 2006. IUCN Red List of Threatened animals. IUCN, Gland, Switzerland.

Ishtiaq, F., L. Guillaumot, S.M. Clegg, A.B. Phillimore, R.A. Black, I.P.F. Owens, N.I. Mundy, and B.C. Sheldon. 2008. Avian haematozoan parasites and their associations with mosquitoes across Southwest Pacific

Islands. Mol. Ecol. 17: 4545-4555.Kim, K.S., Y. Tsuda, and A. Yamada. 2009. Bloodmeal

identification and detection of avian malaria parasite from mosquitoes (Diptera: Culicidae) inhabiting coastal areas of Tokyo Bay, Japan. J. Med. Entomol. 46:1230-1234.

Kimura, M., A.A. Dhondt, and I.J. Lovette. 2006. Phylogeographic structuring of Plasmodium lineages across the North American range of the house finch (Carpodacus mexicanus). J. Parasitol. 92: 1043-1049.

Krzywinski, J., R.C. Wilkerson, and N.J. Besansky. 2001. Evolution of mitochondrial and ribosomal gene sequences in Anophelinae (Diptera: Culicidae): Implications for phylogeny reconstruction. Mol. Phylogenet. Evol. 18: 479-487.

Labarthe, N., M.L. Serrão, Y.F. Melo, S.J. Oliveira, and R. Lourenço-de-Oliveira. 1998. Potential vectors of Dirofilaria immitis (Leidy, 1856) in Itacoatiara, oceanic region of Niterói municipality, State of Rio de Janeiro, Brazil. Memórias do Instituto Oswaldo Cruz 93: 425-432.

Levin, G.A. and R. Moran. 1989. The vascular flora of Isla Socorro, México. San Diego Soc. Nat. Hist. Mem. 16: 1-72.

Levin, II, D.C. Outlaw, F.H. Vargas, and P.G. Parker. 2009. Plasmodium blood parasite found in endangered Galapagos penguins (Spheniscus mendiculus). Biol. Conserv. 142: 3191-3195.

Martínez-Gómez J.E. and R.L. Curry 1996. The conservation statusof the Socorro Mockingbird Mimodes graysoni in 1993-1994. Bird. Conserv. Int. 6:271-283.

Martínez-Gómez J.E., A. Flores-Palacios and R.L. Curry. 2001. Habitat requirements of the Socorro Mockingbird Mimodes graysoni. Ibis143: 456-467.

Martínez-Gómez, J.E., H.M. Horblit, S.G. Stadler, and P.W. Shannon. 2010. Re-introduction of the Socorro Dove, Socorro Island, Revillagigedo Archipelago, Mexico. In: (P.S. Soorae, ed.) Global Re-introduction Perspectives. The IUCN/SSC Re-introduction specialist Group (RSG). Volume II.

Martínez-Gómez, J.E. and J.K. Jacobsen 2004. The conservation statusof Townsend´s Shearwater Puffinus auricularis auricularis. Biol. Conserv. :35-47.

Nayar, J. K., L. Rosen, and J.W. Knight. 1986. Experimental vertical transmission of Saint Louis encephalitis virus by Florida mosquitoes. Am. J. Trop. Med. Hyg. 35: 1296-1301.

Nielsen, E.T. and A.T. Nielsen. 1953. Field observations on the habits of Aedes taeniorhynchus. Ecology 34: 141-156.

Njabo, K.Y., A.J. Cornel, R.N.M. Sehgal, C. Loiseau, W. Buermann, R.J. Harrigan, J. Pollinger, G. Valkiūnas, and T.B. Smith. 2009. Coquillettidia (Culicidae, Diptera) mosquitoes are natural vectors of avian malaria in Africa. Malaria J. 8: 193.

Nylander, J. A. A. 2004. MrModeltest v2. Evolutionary Biology Centre, Uppsala University.

Padilla, L.R., D. Santiago-Alarcón, J. Merkel, R.E. Miller,

Page 8: Implications of Plasmodium parasite infected mosquitoes on ...user et al. 2011 Jour… · Mosquito identification Voucher female and all male mosquitoes were pinned and identified

220 Journal of Vector Ecology June 2011

and P.G. Parker. 2004. Survey for Haemoproteus spp., Trichomonas gallinae, Chlamydophila psittaci, and Salmonella spp. in Galapagos Islands columbiformes. J. Zoo Wildlife Med. 35: 60-64.

Parker, P.G., N.K. Whiteman, and R.E. Miller. 2006. Conservation medicine on the Galápagos Islands: partnerships among behavioral, population, and veterinary scientists. Auk 123: 625-638.

Peck, S. B., J. Heraty, B. Landry, and B.J. Sinclair. 1998. Introduced insect fauna of an oceanic archipelago: the Galápagos Islands, Ecuador. Am. Entomol. 44: 218-237.

Provost, M.W. 1951. The occurrence of salt marsh mosquitoes in the interior of Florida. Fla. Entomol. 34: 48-53.

Ricklefs, R.E. and S.M. Fallon. 2002. Diversification and host switching in avian malaria parasites. Proc. R. Soc. B 269: 885-892.

Ricketts, T.H., E. Dinerstein, T. Boucher, T.M. Brooks, S.H.M. Butchart, M. Hoffmann, J. F. Lamoreux, et al. 2005. Pinpointing and preventing imminent extinctions. Proc. Natl. Acad. Sci. USA 102:18497 -18501.

Santiago-Alarcón, D., D.C. Outlaw, R.E. Ricklefs, and P.G. Parker. 2010. Phylogenetic relationships of haemosporidian parasites in New World Columbiformes, with emphasis on the endemic Galápagos dove. Int. J. Parasitol. 40: 463-470.

SEDESOL. 1994. Decreto por el que se declara como área natural protegida con el carácter de reserva de la Biosfera, a la región conocida como Archipiélago de Revillagigedo, integrada por cuatro áreas: Isla San Benedicto, Isla Clarion o Santa Rosa, Isla Socorro o Santo Tomas e Isla Roca Partida. Diario Oficial de la Federación, 6 June. 4. 45-51.

Smith, K.F., D.F. Sax, and K.D. Lafferty. 2006. Evidence for the role of infectious disease in species extinction and endangerment. Conserv. Biol. 20: 1349-1357.

Snell, H.L., A. Tye, C.E. Causton, and R. Bensted-Smith. 2002. Current status of and threats to the terrestrial biodiversity of Galápagos. In: A Biodiversity Vision for the Galapagos Islands. Charles Darwin Foundation and WWF. Puerto Ayora. pp. 30-47.

Tompkins, D. M. and D. M. Gleeson. 2006. Relationship between avian malaria distribution and an exotic invasive mosquito in New Zealand. J. R. Soc. New Zealand 36: 51-62.

Tompkins, D. M. and R. Poulin. 2006. Parasites and biological invasions. Ecol. Stud. 186: 67-84.

Turell, M. J., M.L. OíGuinn, D.J. Dohm, and J.W. Jones. 2001. Vector competence of North American mosquitoes (Diptera: Culicidae) for West Nile virus. J. Med. Entomol. 38: 130-134.

Valkiūnas, G. 2005. Avian Malaria Parasites and other Haemosporidia. CRC Press, Boca Raton, Florida.

Van Riper, C., S.G. van Riper, M.L. Goff, and M. Laird. 1986. The epizootiology and ecological significance of malaria in Hawaiian land birds. Ecol. Monogr. 56: 327-344.

Vázquez, G. L. 1960. Artrópodos. La Isla Socorro. Monogr. Inst. Geofís. Univ. Nac. Autón. Méx. 2: 126-152 .

Waldenström, J., S. Bensch, D. Hasselquist, and O. Ostman. 2004. A new nested polymerase chain reaction method very efficient in detecting Plasmodium and Haemoproteus infections from avian blood. J. Parasitol. 90: 191-194.

Warner, R.E. 1968. The role of introduced diseases in the extinction of the endemic Hawaiian avifauna. Condor 70: 101-120.

Whiteman, N.K., S.J. Goodman, B.J. Sinclair, T. Walsh, A.A. Cunningham, L.D. Kramer, and P.G. Parker. 2005. Establishment of the avian disease vector Culex quinquefasciatus Say, 1823 (Diptera: Culicidae) on the Galápagos Islands, Ecuador. Ibis 147: 844-847.

Wikelski, M., J. Foufopoulos, H. Vargas, and H. Snell. 2004. Galápagos birds and diseases: Invasive pathogens as threats for island species. Ecol. Soc. 9: 5.